• Aucun résultat trouvé

Navier-Stokes systems with quasimonotone viscosity tensor.

N/A
N/A
Protected

Academic year: 2022

Partager "Navier-Stokes systems with quasimonotone viscosity tensor."

Copied!
17
0
0

Texte intégral

(1)

NAVIER-STOKES SYSTEMS WITH QUASIMONOTONE VISCOSITY TENSOR

Pierre Dreyfuss, Norbert Hungerb¨uhler2

1Mathematics Department, University of Fribourg, P´erolles, CH-1700 Fribourg, SWITZERLAND

e-mail: norbert.hungerbuehler@unifr.ch

2Mathematics Department, University of Fribourg, P´erolles, CH-1700 Fribourg, SWITZERLAND

e-mail: pierre.dreyfuss@unifr.ch

Abstract: In [1] we investigated a class of Navier-Stokes systems which is motivated by models for electrorheological fluids. We obtained an existence result for a weak solution under mild monotonicity assumptions for the viscos- ity tensor. In this article, we continue the analysis of such systems, but with various notions of quasimonotonicity instead of classical pointwise monotonic- ity assumptions. Moreover we allow the external force to be of a more general form.

AMS Subj. Classification: 35Q30, 76D05

Key Words: Navier-Stokes systems, weak and quasimonotonicity conditions, electrorheological fluids

1 Introduction

1.1 Retrospect of former results

In this paragraph we introduce some notations, and we recall the main results established in [1] in order to relate them later on with the new results which we derive below.

Let Ω⊂IRnbe a bounded open domain with Lipschitz boundary. In [1] we considered the following Navier-Stokes system for the velocityu: Ω×[0, T)→ IRn and the pressure P : Ω×[0, T)→IR

This work is supported by the Swiss National Science Foundation under the Grant Number 200020-100051/1

(2)

∂u

∂t −divσ(x, t, u, Du) + (u· ∇)u=f−gradP on Ω×(0, T) (1)

divu= 0 on Ω×(0, T) (2)

u= 0 on ∂Ω×(0, T) (3)

u(·,0) =u0 on Ω (4)

Here, f ∈Lp0(0, T;V0) for somep∈[1 +n+22n ,∞), whereV consists of all func- tions inW01,p(Ω,IRn) with vanishing divergence. Moreoveru0 ∈L2(Ω; IRn) is an arbitrary initial condition satisfying divu0 = 0, and σ satisfies the conditions (NS0)–(NS2) below. We allow the viscosity tensor σ to depend (non-linearly) on x, t, u and Du.

The problem (1)–(4) with theu-dependence of σis motivated by the study of electrorheological fluid flows, as explained in the introduction of [1].

To fix some notation, let IIMm×n denote the real vector space of m ×n matrices equipped with the inner product M : N = MijNij (with the usual summation convention).

The following main assumptions are imposed on the viscosity tensor σ:

(NS0) (Continuity) σ : Ω×(0, T)×IRn×IIMn×n → IIMn×n is a Carath´eodory function, i.e. (x, t)7→σ(x, t, u, F) is measurable for every (u, F)∈IRn× IIMn×n and (u, F) 7→ σ(x, t, u, F) is continuous for almost every (x, t) ∈ Ω×(0, T).

(NS1) (Growth and coercivity) There existc1 >0,c2 >0,λ1 ∈Lp0(Ω×(0, T)), λ2 ∈L1(Ω×(0, T)),λ3 ∈L(p/α)0(Ω×(0, T)), 0< α < p, such that

|σ(x, t, u, F)| 6 λ1(x, t) +c1(|u|p−1 +|F|p−1) σ(x, t, u, F) :F > −λ2(x, t)−λ3(x, t)|u|α+c2|F|p (NS2) (Monotonicity) σ satisfies one of the following conditions:

(a) For all (x, t)∈Ω×(0, T) and all u∈IRn, the map F 7→σ(x, t, u, F) is a C1-function and is monotone, i.e.

(σ(x, t, u, F)−σ(x, t, u, G)) : (F −G)>0 for all (x, t)∈Ω×(0, T),u∈IRm and F, G∈IIMn×n.

(b) There exists a function W : Ω×(0, T)×IRn×IIMn×n → IR such that σ(x, t, u, F) = ∂W∂F (x, t, u, F), and F 7→W(x, t, u, F) is convex and C1 for all (x, t)∈Ω×(0, T) and all u∈IRn.

(c) σ is strictly monotone, i.e. σ is monotone and (σ(x, t, u, F) − σ(x, t, u, G)) : (F −G) = 0 implies F =G.

(3)

We recall that the main point is that, in (a) and (b), it is not required thatσ is strictly monotone or monotone in the variables (u, F) as it is usually assumed in previous work.

We will work with the following function spaces: Let V :={ϕ ∈C0(Ω; IRn) : divϕ= 0}.

Then,V denotes the closure ofV in the spaceW1,p(Ω; IRn). A classical result of de Rham shows, that this space is

V ={ϕ ∈W01,p(Ω; IRn) : divϕ= 0}.

In addition, we will have to work with Ws,2(Ω; IRn), where s >1 + n2. Then, we denote by

Vs := the closure ofV in the space Ws,2(Ω) and

Hq := the closure ofV in the spaceLq(Ω), and H :=H2.

Furthermore, let W denote the space defined by

W :={v ∈Lp(0, T;V) :∂tv ∈Lp0(0, T;V0)},

where the integrals are to be understood in the sense of Bochner and the time- derivative means here the vectorial distributional derivative. We recall that W is continuously embedded in C0([0, T];H) and we always identifyv ∈W with its representative in C0([0, T];H).

The main result we have proved in [1] is the following:

Theorem 1 Assume that σ satisfies the conditions (NS0)–(NS2) for some p ∈ [1 + n+22n ,∞). Then for every f ∈ Lp0(0, T; V0) and every u0 ∈ H, the Navier-Stokes system (1)–(4) has a weak solution (u, P), withu∈ W , in the following sense: For every v ∈Lp(0, T;V) there holds

Z T

0

h∂tu, vidt+ Z T

0

Z

σ(x, t, u, Du) :Dv dx dt+

+ Z T

0

Z

(u· ∇)u·v dx dt= Z T

0

hf, vidt, u(0,·) = u0.

(4)

The weak solution in Theorem 1 is more than a classical weak solution and in particular the energy equality is satisfied (see Remark in [1, p. 245] for more details).

For the proof we used a Faedo-Galerkin method. By using the assumptions in (NS1) we constructed a Galerkin sequence (um) of approximating solutions.

Several compactness properties were then established in [1] which allowed to extract a subsequence um converging weakly to some u in Lp(0, T;V). But then, the monotonicity assumptions (NS2) (a) or (b) do not allow to use the classical monotonicity method (like in [9]) in order to pass to the limit in the Galerkin equations. To overcome this difficulty we then used a refinement of the monotone operator method (inspired by [3]) which involves the theory of Young measures. To this end, we studied the Young measure ν(x,t) generated by the sequence of gradients (Dum), and obtained a div-curl inequality which was the key ingredient to pass to the limit in the Galerkin equation. This inequality was formulated as follows:

Lemma 2 (A div-curl inequality) The Young measureν(x,t) generated by the gradients Dum of the Galerkin approximations um has the property, that for all s ∈[0, T]:

Z s

0

Z

Z

IIMn×n

σ(x, t, u, λ)−σ(x, t, u, Du)

: λ−Du

(x,t)(λ)dxdt60. (5) In a final step we have shown that the existence result in Theorem 1 follows from the div-curl inequality whenever we use one of the monotonicity condi- tions in (NS2). Moreover we have derived some additional properties of the Galerkin approximations: In case (NS2) (a) there holds σ(x, t, um, Dum) * σ(x, t, u, Du) in Lp0(Ω×(0, T)) (for a subsequence), in case (b) we have in addition σ(x, t, um, Dum)→σ(x, t, u, Du) inLβ(Ω×(0, T)), for allβ ∈[1, p0), and in case (c), we even have Dum →Duin Lα(Ω×(0, T)) for all α∈[1, p).

1.2 Extension of the results to quasimonotone viscosity tensors In this paper we intend to consider quasimonotonicity assumptions forσrather than the classical pointwise monotonicity: Instead of (NS2) (a), (b) or (c) which are pointwise monotonicity conditions, we consider the assumptions (NS2) (d) and (e) below which represent monotonicity in an integrated form.

Moreover we consider equation (1) with a source term which is allowed to be of a more general form. More precisely, we replace f ∈ Lp0(0, T;V0) by f satisfying the assumption (Hf):

(Hf) f : Ω×(0, T)×Rn ×Mn×n → Rn is a Carath´eodory function in the sense (NS0). Moreover we assume that one of the following additional conditions hold:

(5)

(i) For a constant β < p−1 and a function λ4 ∈Lp0(Ω×(0, T)) there holds

|f(x, t, u, F)|6λ4(x, t) +C |u|β+|F|β .

(ii) In addition to (i), the function f is independent of the fourth vari- able, or, for a.e. (x, t) ∈ Ω×(0, T) and all u ∈ Rn, the mapping F →f(x, t, u, F) is linear.

We now introduce the definitions of quasimonotonicity which we are going to use:

Definition 3 A functionη: IIMn×n →IIMn×nis calledstrictly quasimonotone, if there exist constants c >0and r > 0such that

Z

(η(Du)−η(Dv)) : (Du−Dv)dx>c Z

|Du−Dv|rdx for all u, v ∈W01,p(Ω).

We say that η is strictly p-quasimonotone, if Z

IIMn×n

(η(λ)−η(¯λ)) : (λ−λ)dν(λ)¯ >0

for all homogeneous W1,p gradient Young measures ν with center of mass λ¯=hν,idi which are not a single Dirac mass.

Remarks: (a) Note that the notion of p-quasimonotonicity is phrased in terms of gradient Young measures. Notice that although quasimonotonicity is “monotonicity in integrated form”, the gradient of a quasiconvex function is not necessarily strictly p-quasimonotone. A simple example of a strictly p-quasimonotone function is the following: Assume that ηsatisfies the growth condition

|η(F)|6C|F|p−1 with p > 1 and the structure condition

Z

(η(F +Dϕ)−η(F)) :Dϕ dx>c Z

|Dϕ|pdx

for a constant c > 0 and for all ϕ ∈ C0(Ω) and all F ∈ IIMn×n. Then η is strictly p-quasimonotone. This follows easily from the definition if one uses that for every W1,p gradient Young measure ν there exists a sequence (Dvk) generating ν for which (|Dvk|p) is equiintegrable (see [5], [7]).

(b) In [10], Zhang introduced the following notion of quasimonotonicity:

The continuous function η : IIMN×n → IIMN×n is quasimonotone (in the sense

(6)

of Zhang) if for every F ∈ IIMN×n, every open subset G of IRn and every ϕ ∈C01(G; IRN) there holds

Z

G

η(F +Dϕ) :Dϕdx>0.

However, he uses a stronger notion of quasimonotonicity to prove his results, namely, that

Z

G

η(F +Dϕ) :Dϕdx>c Z

G

|Dϕ|pdx (6) for a fixed constant c >0, along with the growth condition

|η(F)|6C|F|p−1 (7) for some constantC > 0. We would like to indicate that our definition of strict p-quasimonotonicity can be considered as a generalization of Zhang’s notion of quasimonotonicity. In fact, a function which is quasimonotone in the sense of Zhang, i.e. which satisfies (6) and (7), is strictly p-quasimonotone. To see this, we consider a homogeneous W1,p gradient Young measure ν with center of mass ¯λ:=hν,idiand which is not a single Dirac mass. Then (for some fixed domain G), there exists a sequence (ϕk) in C0(G) such that the sequence of gradients (Dϕk) generates ν and such that (|Dϕk|) is equiintegrable (see Remark (a)). By (6) there holds

Z

G

η(F +Dϕk) :Dϕkdx>c Z

G

|Dϕk|pdx. (8) The sequence (|Dϕk|) is equiintegrable and, by (7), (η(F +Dϕk) : Dϕk) is also equiintegrable. Hence by Ball’s fundamental theorem on Young measures (see [2]), we obtain from (8), that

Z

G

Z

IIMn×n

η(F +λ) :λdνx(λ)dx>c Z

G

Z

IIMn×n

|λ|px(λ)dx, and since ν, by hypothesis, is homogeneous,

Z

IIMn×n

η(F +λ) :λdν(λ)>c Z

IIMn×n

|λ|pdν(λ).

A substitution λ=λ0−λ, and the special choice¯ F = ¯λ yields Z

IIMn×n

η(λ0) : (λ0−λ)dν(λ¯ 0)>c Z

IIMn×n

0−¯λ|pdν(λ0)>0 (9) since ν is not a single Dirac mass. Moreover,

Z

IIMn×n

η(¯λ) : (λ0−¯λ)dν(λ0) =η(¯λ) : Z

IIMn×n

λ0dν(λ0)

| {z }

λ

−η(¯λ) : ¯λ Z

IIMn×n

dν(λ0)

| {z }

=1

= 0.

(10)

(7)

If we subtract (10) from (9), we arrive at Z

IIMn×n

(η(λ0)−η(¯λ)) : (λ0−λ)dν(λ¯ 0)>0,

and hence η is strictly p-quasimonotone. We end this remark by noting that we can replace the power p on the right hand side of (6) by any power r > 0

and still have the same conclusion. ♦

So, in addition to (NS2) (a), (b) and (c), we will now consider the condi- tions:

(NS2) (Monotonicity) σ satisfies one of the following conditions:

(d) for a.e (x, t) ∈ Ω × (0, T) and all u ∈ Rn, the function F → σ(x, t, u, F) is strictly quasimonotone.

(e) for a.e (x, t) ∈ Ω × (0, T) and all u ∈ Rn, the function F → σ(x, t, u, F) is strictlyp-quasimonotone.

The main result we prove in this paper is the following:

Theorem 4 Assume thatσsatisfies the conditions (NS0) and (NS1) for some p∈[1 + n+22n ,∞). Let u0 be given in H. Then we have:

(i) If σ satisfies one of the condition (NS2) (c), (d) or (e) then for every f satisfying (Hf) (i), the Navier-Stokes system (1)–(4) has a weak solution (u, P), with u∈W .

(ii) If σ satisfies one of the condition (NS2) (a) or (b) then for every f satisfying (Hf) (ii) the same conclusion holds.

Remark: The notion of a solution u ∈ W in Theorem 4 is the same as in Theorem 1: We have u(0,·) =u0 and there holds

Z T

0

h∂tu, vidt+ Z T

0

Z

σ(x, t, u, Du) :Dv dx dt+ Z T

0

Z

(u· ∇)u·v dx dt=

= Z T

0

Z

f(x, t, u, Du)·v dx dt ∀v ∈Lp(0, T;V).

1.3 Organization of the paper

In Section 2 we will resume the relevant results presented in [1, Section 1–

6]. We will prove first that for every f satisfying (Hf) (i) we can construct a Galerkin sequence um of approximating solutions. In a second step we will show that the convergence properties for um established in [1] remain valid.

We will then recover the same properties for the Young measure associated to (um, Dum), and actually the div-curl inequality holds again.

(8)

In Section 3 we pass to the limit m → ∞ in the Galerkin equations and prove Theorem 4. Like in [1], the key ingredient for identifying the weak limit σ(x, t, um, Dum) * σ(x, t, u, Du) (where u is the weak limit in Lp(0, T;V) of a relabeled subsequence of um) is the div-curl lemma (unless for the easier situation when (NS2) (d) holds). In the cases (NS2) (c), (d) or (e) we also get that Dum → Du in measure, which allows to conclude the first part of Theorem 4. In the situation (NS2) (a) or (b), this additional property of convergence does not hold in general and we have to consider the stronger assumption (Hf) (ii) instead of (Hf) (i) in order to identify the weak limit in f(x, t, um, Dum)* f(x, t, u, Du).

2 The Galerkin approximation 2.1 The Galerkin base

Let the functions wi ∈ Vs be a Galerkin base, as introduced in [1, Section 2]. We have shown that W := {w1, w2, . . .} is an orthonormal Hilbert base of H. In particular, the L2-orthogonal projector Pm : H → H onto span(w1, w2, . . . , wm),m ∈Nis defined by the formula

Pmu=

m

X

i=1

(wi, u)Hwi. (11)

Of course, the operator norm kPmkL (H,H) = 1. We have also shown that kPmkL(Vs,Vs) = 1 and thatPmconverges pointwise to the identity inL (Vs, Vs).

2.2 The Galerkin approximation

Letm ∈N. In the (Faedo-)Galerkin method one makes an Ansatz for approx- imating solutions um of the form

um(x, t) =

m

X

i=1

cmi(t)wi(x), (12)

where cmi : [0, T) → IR are supposed to be continuous bounded functions.

We take care of the initial condition (4) by choosing the initial coefficients cmi:=cmi(0) = (u0, wi)L2 such that

um(·,0) =

m

X

i=1

cmiwi(·)→u0 in L2(Ω) as m→ ∞. (13)

(9)

We try to determine the coefficientscmi(t) in such a way, that for everym ∈N the system of ordinary differential equations

(∂tum, wj)H + Z

σ(x, t, um, Dum) :Dwjdx+b(um, um, wj) =

= Z

f(x, t, um, Dum)·wjdx (14) (with j ∈ {1,2, . . . , m}) is satisfied in the sense of distributions. In (14), we used the shorthand notation

b(u, v, w) :=

Z

((u· ∇)v)·w dx.

Let ε, J, r and K be the quantities introduced in [1, Section 3.1]. For any j = 1, . . . , m we can verify (by using the assumption (Hf)) that the function Θ :J×K →R defined by

Θ(t, c1, . . . , cm) :=

Z

f(x, t,

m

X

i=1

ciwi,

m

X

i=1

ciDwi)·wjdx is a Carath´eodory function. Moreover we obtain the estimate:

|Θ(t, c1, . . . , cm)|6C Z

λ4(x, t)dx+C,

where C may depend on m and r but is independent of t. These results allow to conclude as in [1, Section 3.1] that there exists a local solution for equation (14). This solution um is on the form (12) and it verifies (14) in the sense D 0(0, ε0), for some ε0 >0 which, for the moment, may depend on m.

This local solution can be extended to the whole interval [0, T) independent ofm. To do this one can use the arguments presented in [1, Section 3.2]. More precisely, let τ be arbitrary in the existence interval. We have to replace the term III =Rτ

0hf(t), umidtbyIII =Rτ 0

R

f(x, t, um, Dum)·umdxdt. By using the growth condition (Hf) (i) we then obtain

III 6kλ4kLp0

(Ω×(0,T))kumkLp(Ω×(0,T))+Ckumkβ+1Lp(0,T;V).

This inequality, in combination with the estimates for the terms I andII from [1, Section 3.2] which remain unchanged, permits again to obtain the estimate

|(cmi(τ))i=1,...,m|2IRm =kum(·, τ)k2L2(Ω)6C¯

for a constant ¯C which is independent of τ (and of m). It follows that the functions cmj can be extended to the whole interval [0, T] and um(x, t) = Pm

j=1cmj(t)wj(x) is a solution (not necessarily unique) of (14) in the sense D 0(0, T). Moreover we obtain again

kumkC0([0,T];L2(Ω)) 6C. (15)

(10)

2.3 Basic convergence properties

We easily recover the basic convergence properties presented in [1, Section 4.1], i.e. we may extract a subsequence, still denoted by um, verifying

um * u in L(0, T;H) (16) um * u in Lp(0, T;V) (17)

−divσ(x, t, um, Dum)* χ in Lp0(0, T;V0) (18) for some u ∈ Lp(0, T;V)∩L(0, T;H) and χ ∈ Lp0(0, T;V0). Moreover, by using (17) together with (Hf) (i) we may also assume that

f(x, t, um.Dum)* ξ in Lp0(Ω×(0, T)) (19) for some ξ ∈Lp0(Ω×(0, T)).

The principal difficulty will be, as in [1], to show thatχ=−divσ(x, t, u, Du).

Moreover here, we will also have to identify ξ with f(x, t, u, Du).

2.4 Convergence in measure

We recall first that for anyqsatisfying 2< q < p := n−pnp we have the following chain of continuous injections:

V ,→i Hq i0

,→H

γ

=H0 ,→i1 Vs0. (20) Here, H ∼= H0 is the canonical isomorphism γ between the Hilbert space H and its dual. We take over from [1] the notation j ◦i to denote the canonical injection of V into Vs0, that is, for u∈V :

hj◦i◦u, vi= Z

uv dx ∀v ∈Vs.

Next, we consider the time derivative of um inD 0(0, T;Vs0), which is in fact a function in Lp0(0, T;Vs0) given by the formula :

h∂t(j◦i◦um), vi= Z

tum(x, t)Pmv(x)dx=− Z

σ(x, t, um, Dum) :D(Pmv)dx−

−b(um, um, Pmv) + Z

f(x, t, um, Dum)Pmvdx ∀v ∈Vs, a.e. t ∈(0, T).

(21) By using the fact that um is a bounded sequence inLp(0, T;V) together with the growth property (Hf) (i) we obtain

Z

f(x, t, um, Dum).Pmvdx

6CkvkVskf(x, t, um, Dum)kL1(Ω)m(t)kvkV s,

(11)

where (γm) is a bounded sequence in Lp0(0, T).

Consequently we obtain :

|h∂t(j◦i◦um), vi|6Cm(t)kvkVs (22) where (Cm) is a bounded sequence in Lp0(0, T).

Remark: The estimate (25) in [1, p. 255] needs to be modified in the following way: C ∈Lp0(0, T) should be replaced by (Cm), a bounded sequence in Lp0(0, T), like here in estimate (22). This modification has no consequence on the results presented as we will see in the sequel.

From (22) we conclude indeed, that {∂tj◦i◦um}m is a bounded sequence in Lp0(0, T;Vs0).

Consequently, by [1, Lemma 2], we may assume (for a further subsequence) um →u in Lp(0, T;Lq(Ω)) for all q < p and in measure on Ω×(0, T).

(23) 2.5 A regularity result for u

The arguments developed in [1, Section 4.3] are easily carried over: It follows that the time derivative ∂t(j ◦i◦u) is given by the formula

h∂t(j◦i◦u), vi= Z

ξvdx− hχ, vi −b(u, u, v), ∀v ∈Vs, a.e. t ∈(0, T). (24) This permits again to obtain the property ∂t(j ◦i◦u) ∈ Lp0(0, T;V0), which implies that u∈W .

2.6 The limiting time values for u

We remark first that by the boundedness of the sequence ∂t(j ◦ i◦ um) in Lp0(0, T;Vs0) we may extract a subsequence (not relabeled) such that

t(j◦i◦um)* ∂t(j ◦i◦u) inLp0(0, T;Vs0). (25) Next, by the same manner as in [1, Section 4.4], we get

um(x,0)→u0(x) =u(x,0) inH, (26) um(·, T)* u(·, T), inH. (27) The property (27) can be extended to all t∈[0, T]. Indeed, we have:

Lemma 5 There exists a subsequence of {um} (still denoted byum) with the property that for every t∈[0, T]

um(·, t)* u(·, t) in H. (28)

By (23) the convergence in (28) is actually strong for almost all t∈[0, T].

(12)

Proof. We can take over the proof of [1, Lemma 4] with taking into account the remark after the relation (22). The estimate (25) in [1, p. 255] is not true and we have to replace C(t) by Cm(t), where Cm(t) is a bounded sequence in Lp0(0, T). In fact, after this correction the relation (25) in [1] is the same estimate as (22) in this paper. Recall next that we consider p < ∞ and thus the boundedness ofCm inLp0(0, T) implies the equiintegrability property.

Consequently the estimate (43) in [1] holds true when we replaceC(t) byCm(t)

and the rest of the proof follows. 2

2.7 The Young measure generated by the Galerkin approximation The sequence (or at least a subsequence) of the gradients Dum generates a Young measure ν(x,t), and since um converges in measure to u on Ω×(0, T), the sequence (um, Dum) generates the Young measure δu(x,t)⊗ν(x,t) (see, e.g., [6]). Now, we collect some facts about the Young measure ν in the following proposition:

Proposition 6 The Young measureν(x,t)generated by the sequence {Dum}m has the following properties:

(i) ν(x,t) is a probability measure on IIMn×nfor almost all (x, t)∈Ω×(0, T).

(ii) ν(x,t) satisfiesDu(x, t) = hν(x,t),idi for almost every(x, t)∈Ω×(0, T).

(iii) ν(x,t) has finite p-th moment for almost all (x, t)∈Ω×(0, T).

(iv) ν(x,t) is a homogeneous W1,p gradient Young measure for almost all (x, t)∈Ω×(0, T).

Proof. For (i), (ii) and (iii) see the proof of Proposition 5 in [1].

(iv) We have to show, that {ν(x,t)}x∈Ω is for almost all t ∈ (0, T) a W1,p gradient Young measure. To see this, we take a quasiconvex function q on IIMn×n with q(F)/|F| → 1 as F → ∞. Then, we fix x ∈ Ω, δ ∈ (0,1) and use inequality (1.21) from [8, Lemma 1.6] with u replaced by um(x, t), with a:=u(x, t)−Du(x, t)xand withX :=Du(x, t). Furthermore, we chooser >0 such that Br(x) ⊂ Ω. Observe, that the singular part of the distributional gradient vanishes for um and, after integrating the inequality over the time

(13)

interval [t0−ε, t0+ε]⊂(0, T), we get Z t0

t0−ε

Z

Br(x)

q(Dum(y, t))dydt+

+ 1

(1−δ)r Z t0

t0−ε

Z

Br(x)\Bδr(x)

|um(y, t)−u(x, t)−Du(x, t)(y−x)|dydt>

>|Bδr(x)|

Z t0 t0−ε

q(Du(x, t))dt.

Letting m tend to infinity in the inequality above, we obtain Z t0

t0−ε

Z

Br(x)

Z

IIMn×n

q(λ)dν(y,t)(λ)dydt+

+ 1

(1−δ)r Z t0

t0−ε

Z

Br(x)\Bδr(x)

|u(y, t)−u(x, t) +Du(x, t)(y−x)|dydt>

>|Bδr(x)|

Z t0 t0−ε

q(Du(x, t))dt.

Now, we letε →0 andr→0 and use the differentiability properties of Sobolev functions (see, e.g., [4]) and obtain, that for almost all (x, t0)∈Ω×(0, T)

Z

IIMn×n

q(λ)dν(x,t0)(λ)> |Bδr(x)|

|Br(x)| q(Du(x, t0)).

Since δ ∈ (0,1) was arbitrary, we conclude that Jensen’s inequality holds true for q and the measure ν(x,t) for almost all (x, t) ∈ Ω ×(0, T). Using the characterization of W1,p gradient Young measures of [7] (e.g., in the form of [8, Theorem 8.1]), we conclude that in fact {νx,t}x∈Ω is a W1,p gradient Young measure on Ω for almost all t∈(0, T). By the localization principle for gradient Young measures, we conclude then, thatν(x,t) is a homogeneous W1,p gradient Young measure for almost all (x, t)∈Ω×(0, T). 2

2.8 A Navier-Stokes div-curl inequality

In this section, we prove a Navier-Stokes version of a “div-curl Lemma” which will be the key ingredient to obtain χ =−divσ(x, t, u, Du). The results and the arguments presented in [1, Section 6] can be carried over with only minor modifications which we want to explain in the sequel.

In a first step we note that Z s2

s1

hχ, uidt+1

2ku(·, s2)k2H = Z s2

s1

Z

ξ·udxdt+1

2ku(·, s1)k2H, ∀06s1 6s2 6T.

(29)

(14)

This follows in the same way as the energy equality (46) in [1].

Next, we establish the following lemma:

Lemma 7 (A div-curl inequality) The Young measureν(x,t) generated by the gradients Dum of the Galerkin approximations um has the property that for all s ∈[0, T] :

Z s

0

Z

Z

IIMn×n

σ(x, t, u, λ)−σ(x, t, u, Du)

: λ−Du

(x,t)(λ)dxdt60. (30) Proof. We follow the calculations in [1, Section 6.2]. The inequality (48) in [1] can be obtained without any change, and inequality (49) holds true if we replace hf, ui by R

ξudx. By using the property (19) together with (23) we then obtain that

m→∞lim Z s

0

Z

f(x, t, um, Dum)·umdxdt= Z s

0

Z

ξ·udxdt.

Hence, by using (26) with (28) we obtain lim inf

m→∞

Z s

0

Z

σ(x, t, um, Dum) :Dumdxdt 6

Z s

0

Z

ξ·udxdt−1

2ku(·, s)k2H +1

2ku(·,0)k2H.

The rest of the proof is identical to the proof in [1]. 2 Remarks:

(i) In the proof of the div-curl lemma of we do not use any monotonicity assumption.

(ii) An intermediary result is that, for all s∈[0, T], there holds lim inf

m→∞

Z s

0

Z

(σ(x, t, um, Dum)−σ(x, t, u, Du)) : (Dum−Du)dxdt60.

To see this, repeat the proof of Lemma 6 in [1] with the modifications indicated above.

3 Passage to the limit

Here we will pass to the limit m → ∞ in the Galerkin equations and prove Theorem 4. The first step is to identify the weak limit σ(x, t, um, Dum) * σ(x, t, u, Du). This will follow from every monotonicity assumption listed in (NS2). In Subsection 3.1, we treat the special cases (NS2) (c), (d) and (e) for which we also obtain the convergenceDum→Du in measure. The conclusion is then given in the last subsection.

(15)

3.1 The cases (NS2) (c), (d) and (e)

In these three cases we will prove that we may extract a subsequence with the property

Dum →Du in measure on Ω×(0, T). (31) We consider first the case (NS2) (d). In this situation, elementary arguments are actually sufficient to prove (31), and we do not need the div-curl inequality:

Observe that we have Z T

0

Z

|Dum−Du|rdxdt6

6C

Z T

0

Z

(σ(x, t, um, Dum)−σ(x, t, um, Du)) : (Dum−Du)dxdt

6C

Z T

0

Z

(σ(x, t, um, Dum)−σ(x, t, u, Du)) : (Dum−Du)dxdt+

+C Z T

0

Z

(σ(x, t, u, Du)−σ(x, t, um, Du)) : (Dum−Du)dxdt.

(32) We remark now that the limit inferior of the first term on the right hand side of (32) is less than or equal to zero (see remark (ii) after lemma 7). The second term vanishes when m tends to infinity because of (23). It follows that

lim inf

m→∞

Z T

0

Z

|Dum−Du|rdxdt= 0, and thus (31) holds for a subsequence.

In case (NS2) (c) we can take over the proof presented in [1]. It remains to consider the case (NS2) (e). We suppose that ν(x,t) is not a Dirac mass on a set (x, t) ∈ M ⊂ Ω×(0, T) of positive Lebesgue measure |M| > 0. Then, by the strict p-quasimonotonicity of σ(x, t, u,·), and the fact that ν(x,t) is a homogeneous W1,p gradient Young measure (see Section 2.7) for almost all (x, t)∈Ω×(0, T), we have for a.e. (x, t)∈M

Z

IIMn×n

σ(x, t, u, λ) :λdν(x,t)(λ)>

>

Z

IIMn×n

σ(x, t, u, λ)dν(x,t)(λ) : Z

IIMn×n

λdν(x,t)(λ)

| {z }

=Du(x, t) .

(16)

Hence, by integrating over Ω×(0, T), we get together with Lemma 7 Z T

0

Z

Z

IIMn×n

σ(x, t, u, λ)dν(x,t)(λ) :Du(x, t)dxdt>

>

Z T

0

Z

Z

IIMn×n

σ(x, t, u, λ) :λdν(x,t)(λ)dxdt >

>

Z T

0

Z

Z

IIMn×n

σ(x, t, u, λ)dν(x,t)(λ) :Du(x, t)dxdt which is a contradiction. Hence, we have ν(x,t) = δDu(x,t) for almost every (x, t) ∈ Ω×(0, T). From this, it follows that Dum → Du on Ω ×(0, T) in measure for m → ∞ (see, e.g., [6]).

3.2 Conclusion For the cases (NS2) (c), (d) and (e) there holds

σ(x, t, um, Dum)→σ(x, t, u, Du) inLβ(Ω×(0, T)),∀β ∈[1, p0), (33) f(x, t, um, Dum)→f(x, t, u, Du) inLβ(Ω×(0, T)),∀β∈[1, p0). (34) To see this, just use (31), the boundedness of the sequences σ(x, t, um, Dum) and f(x, t, um, Dum) inLp0(Ω×(0, T)), and apply the Vitali convergence the- orem. It then follows that

−divσ(x, t, um, Dum)* χ=−divσ(x, t, u, Du) inLp0(0, T;V0), (35) f(x, t, um.Dum)* ξ =f(x, t, u, Du) inLp0(Ω×(0, T)), (36) These properties are sufficient to pass to the limit in the Galerkin equations and to conclude the proof of Theorem 4 in case (i) (see [1, p. 266]).

For the remaining cases (NS2) (a) and (b) the property (31) does not hold in general, but we however obtain σ(x, t, um, Dum) * σ(x, t, u, Du) in Lp0(Ω×(0, T)) by using Lemma 7 in [1, Section 7]. This suffices to conclude the proof of Theorem 4 in case (ii): In fact, in this situation we have assumed that f satisfies the assumption (Hf) (ii), and it remains to show that the property (36) still holds without (31). By using (23) we easily verify that it is true for the particular situation in (Hf) (ii), when f is independent of the fourth variable. In the other situation we have that, for a.e (x, t)∈Ω×(0, T) and all u ∈ Rn, the mapping F → f(x, t, u, F) is linear. Here we argue as follows to identify the weak limit ξ in (36):

f(x, t, um, Dum)* Z

IIMn×n

f(x, t, u, λ)dν(x,t)(λ)

=f(x, t, u,·)◦ Z

IIMn×n

λdν(x,t)(λ) =f(x, t, u, Du),

(17)

where for the last equality we have used the property (ii) of Proposition 6.

We can now again pass to the limit in the Galerkin equations and conclude the proof of Theorem 4 in case (ii). We end this discussion by noting, that the energy equality (29) holds true with ξ replaced by f(x, t, u, Du) and with χ replaced by −divσ(x, t, u, Du).

References

[1] P. Dreyfuss, N. Hungerb¨uhler: Results on a Navier-Stokes system with applications to electrorheological fluid flow. Int. J. of Pure and Appl.

Math. 14, No. 2 (2004), 241–271.

[2] J. M. Ball: A version of the fundamental theorem for Young measures. In:

Partial differential equations and continuum models of phase transitions:

Proceedings of an NSF-CNRS joint seminar. Springer, 1989.

[3] G. Dolzmann, N. Hungerb¨uhler, S. M¨uller: Non-linear elliptic systems with measure-valued right hand side. Math. Z. 226 (1997), 545–574.

[4] L. C. Evans, R. F. Gariepy: Measure theory and fine properties of func- tions. Boca Raton, CRC Press, 1992.

[5] I. Fonseca, S. M¨uller, P. Pedregal: Analysis of concentration and oscilla- tion effects generated by gradients. SIAM J. Math. Anal. 29 (1998), no.

3, 736–756.

[6] N. Hungerb¨uhler: A refinement of Ball’s Theorem on Young measures.

New York J. Math. 3 (1997), 48–53.

[7] D. Kinderlehrer, P. Pedregal: Young measures generated by sequences in Sobolev spaces. J. Geom. Anal. 4, No. 1 (1994), 59–90.

[8] J. Kristensen: Lower semicontinuity in spaces of weakly differentiable functions. Math. Ann. 313 (1999), 653–710.

[9] J.-L. Lions: Quelques m´ethodes de r´esolution des probl`emes aux limites non lin´eaires. Dunod; Gauthier-Villars, Paris 1969.

[10] K.-W. Zhang: On the Dirichlet problem for a class of quasilinear elliptic systems of partial differential equations in divergence form. Partial differ- ential equations (Tianjin, 1986), 262–277, Lecture Notes in Math., 1306, Springer, Berlin-New York, 1988.

Références

Documents relatifs

When considering all the profiles having the same horizontal scale (1 here), the point is therefore to choose the smallest vertical scale (2 n here) and to write the decomposition

Starting from isentropic compressible Navier-Stokes equations with growth term in the con- tinuity equation, we rigorously justify that performing an incompressible limit one arrives

This decomposition is used to prove that the eigenvalues of the Navier–Stokes operator in the inviscid limit converge precisely to the eigenvalues of the Euler operator beyond

In the same way, we obtain here Leray’s weak solutions directly as the incompressible scaling limit of the BGK kinetic model, which is a relaxation model associated with the

In section 3, we study the regularity of solutions to the Stokes equations by using the lifting method in the case of regular data, and by the transposition method in the case of

The aim of this section is to show that we can extract from (u ε , p ε ) ε&gt;0 a subsequence that converges to a turbulent solution of the NSE (1.1) (see Definition 2.3), which

Marius Mitrea, Sylvie Monniaux. The regularity of the Stokes operator and the Fujita–Kato approach to the Navier–Stokes initial value problem in Lipschitz domains.. approach to

Lemma 1.2 also extends to Lemma 2.13 in connection with Theorem 2.9 (critical case s = q N ∗ −1 ), and to Lemma 2.14 (case q = N −1, with a uniform convergence result) in